You are on page 1of 17

This article was downloaded by: [Tulane University]

On: 08 January 2015, At: 06:39


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Phosphorus, Sulfur, and Silicon and the


Related Elements
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gpss20

1,6- And 1,7-Regioisomers of Perylene


Tetracarboxylic Dianhydride and
Diimide: The Effects of Neutral Bay
Substituents on the Electrochemical and
Structural Properties
a a a
Nisha V. Handa , Laura D. Shirtcliff , Barry K. Lavine , Douglas R.
b a
Powell & Darrell K. Berlin
a
Click for updates Department of Chemistry, Oklahoma State University, Stillwater,
OK, USA
b
Department of Chemistry & Biochemistry, University of Oklahoma,
Norman, OK, USA
Accepted author version posted online: 25 Feb 2014.Published
online: 12 Jun 2014.

To cite this article: Nisha V. Handa, Laura D. Shirtcliff, Barry K. Lavine, Douglas R. Powell & Darrell
K. Berlin (2014) 1,6- And 1,7-Regioisomers of Perylene Tetracarboxylic Dianhydride and Diimide: The
Effects of Neutral Bay Substituents on the Electrochemical and Structural Properties, Phosphorus,
Sulfur, and Silicon and the Related Elements, 189:6, 738-752, DOI: 10.1080/10426507.2013.855769

To link to this article: http://dx.doi.org/10.1080/10426507.2013.855769

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Tulane University] at 06:39 08 January 2015
Phosphorus, Sulfur, and Silicon, 189:738–752, 2014
Copyright 
C Taylor & Francis Group, LLC

ISSN: 1042-6507 print / 1563-5325 online


DOI: 10.1080/10426507.2013.855769

1,6- AND 1,7-REGIOISOMERS OF PERYLENE


TETRACARBOXYLIC DIANHYDRIDE AND DIIMIDE: THE
EFFECTS OF NEUTRAL BAY SUBSTITUENTS ON THE
ELECTROCHEMICAL AND STRUCTURAL PROPERTIES

Nisha V. Handa,1 Laura D. Shirtcliff,1 Barry K. Lavine,1


Douglas R. Powell,2 and K. Darrell Berlin1
1
Department of Chemistry, Oklahoma State University, Stillwater, OK 74078, USA
2
Department of Chemistry & Biochemistry, University of Oklahoma, Norman, OK
Downloaded by [Tulane University] at 06:39 08 January 2015

73019, USA
GRAPHICAL ABSTRACT

Abstract The electrochemical and structural properties of a series of 1,6- and 1,7-regioisomers
of different sized bay-appended perylene diimides (PDIs) and perylene tetracarboxylic dianhy-
drides (PTCDs) were assessed. Steric effects by large bay substituents triphenylsilylacetylene
and tritylacetylene play a major role in the geometry of the solid and solution states. New
triphenylsilylacetyene and 1-pentynyl derivatives were prepared and characterized. Suitable
crystals for X-ray analysis of tritylacetylene and n-hexyl compounds illustrated the structural
alterations in the bay region. The bulky tritylacetylene appended PDI assumed a nearly planar
π -configuration, equivalent to an unsubstituted PDI. In contrast, a slender and less bulky hexyl
chain incorporated PDI underwent a significant twisting of the central core of PDI. Neutral
and conjugated groups at the bay region of PDI enhanced its reductive capability. In contrast,
incorporation of neutral and nonconjugated groups at the bay region slightly diminished the
reductive capability of resulting PDI derivative. PTCDs consisting of both bulky and slender
groups were reduced significantly more readily in relation to the respective PDIs. Electro-
chemical reductive properties of selected PDIs and PTCDs were obtained along with optical
properties of 1,6- and 1,7-PDIs.

Keywords Regioisomers of perylene tetracarboxylic dianhydride and diimide; synthesis,


electrochemical and structural properties; bay substituents

Received 5 July 2013; accepted 11 October 2013.


Address correspondence to K. Darrell Berlin, Department of Chemistry, Oklahoma State University Still-
water, Oklahoma, PS I, 422, Stillwater, OK 74078, USA. E-mail: kenneth.d.berlin@okstate.edu
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/gpss.

738
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 739

INTRODUCTION
Perylene diimides (PDIs) have been used as commercial dyes and pigments since
the 1950s due to their high photo and thermal stability, excellent tinctorial strength and
relative chemical inertness.1 In the mid-1990s, PDIs began to be investigated in a variety
of other applications due to their high stability, in addition to their other excellent prop-
erties such as planar π -conjugation, high-fluorescence quantum yields, strong absorption
in the visible region, and good electron mobility.1–4 This combination of properties makes
PDIs attractive materials for both biological and material sciences.2,5–7 These applications
include photodynamic therapy,8 fingerprint detection,9 heparin detection,10,11 artificial pho-
tosynthetic systems,12,13 and a variety of other fluorescent switches and sensors.14 PDIs are
also being explored as n-type semiconductors in molecular devices such as organic field
effect transitions (OFETs),15 organic voltaics (OPVs),16,17 and organic light emitting diodes
(OLEDs).18 Next to fullerenes, PDIs are currently one of the most widely investigated single
molecule n-type organic semiconductors.1
Downloaded by [Tulane University] at 06:39 08 January 2015

For all of their excellent properties, the utility of PDIs in solution-processed or-
ganic devices has been somewhat limited due to their strong propensity to aggregate via
π –π stacking. The parent compound, perylene tetracarboxylic dianhydride (PTCD), is not
amenable to solution-processed organic electronics due to its total insolubility in organic
solvents. However, the stability of PTCD sparked a flurry of synthetic research toward
soluble PDI derivatives. PTCDs can be made soluble in common organic solvents via syn-
thetic modification to the imide, bay (1,6,7,12 positions),1,19–21 and ortho regions (2,5,8,11
positions).22,23 Adding steric bulk at the imide position decreases intermolecular aggrega-
tion of the perylene core by decreasing the propensity for π –π stacking; the greater the
bulk at the imide, the greater the solubility. Substitution at the imide position negligibly
affects the absorption and emission of PDIs due to the presence of nodes in the HOMO
and LUMO orbitals of the imide nitrogen.1,24 The optical and electrochemical properties
of PDIs, however, can be modified considerably through substitution with π -donor and
acceptor groups at the bay region.2 Substitution at the perylene core can, however, also
result in deformation from planarity which causes hypsochromic shifts.2
Standard synthetic literature methods to manipulate PDIs for applications primarily
rely upon increasing solubility by addition of bulky groups at the imide site and fine-tuning
the electronic properties through functionalization of the bay region with electron donor
or acceptor groups. Various synthetic strategies have been investigated to increase the sol-
ubility of the PDI chromophore, with the result of a seminal patent by BASF in 1997
describing the synthesis of 1,7-dibromoperylene-3,4,9,10-tetracrboxylic diimide which,
consequently, can undergo a variety of SN Ar substitution and transition metal catalyzed
cross-coupling reactions.2,25,26 Following disclosure of the patent, several research groups
utilized the procedure to increase the solubility and tune the optical, electrochemical, and
electronic properties through functionalization with electron donor or acceptor groups.27
However, in 2004, it was revealed that 1,7-dibromo-PDIs synthesized by BASF’s proce-
dure actually were composed of a 4:1 regioisomeric mixture of 1,7- and 1,6-dibromo PDIs
rather than the pure 1,7-regioisomer.28 The pure 1,7-regioisomer can only be obtained by
successive recrystallization using CH2 Cl2 :MeOH (3:1) as reported by Wurthner et al.28
The pure 1,6-dibromo-regioisomer, however, cannot be obtained following the aforemen-
tioned procedure. Subsequently, in the last two years, a few reports have been published
in the literature outlining separation of the two regioisomers by column chromatography
in moderate to good yields.29–32 The bulk of the articles highlighting the separation of the
740 NISHA V. HANDA ET AL.

two regioisomers reported to date identify appending bulky groups for successful chro-
matographic purification. This has been attributed to the fact that with bulky substituents,
there is less aggregation, and the regioisomers separate more facilely due to preferential
aggregation.33 We report herein our results on the synthesis and efficient separation of 1,6-
and 1,7-PDIs and PTCDs with both bulky and slender bay substituents. The spectroscopic,
electrochemical, and structural properties of our series of PDIs and PTCDs with a variety
of differently sized functional groups are also reported. We focused on the analysis and
comparison of substitution patterns and size correlations.
This article is an elaboration of our initial communication reporting on the efficient
separation of 1,6- and 1,7-PDIs and PTCDs.34 We have expanded the electronics, shapes,
and sizes of the substituents to yield a library of regioisomers to compare their resulting
properties. The significant effect of neutral conjugated and nonconjugated groups at the
bay region on the electrochemical and structural properties of resulting PDIs and PTCDs
have also been discussed.
Downloaded by [Tulane University] at 06:39 08 January 2015

RESULTS AND DISCUSSION


General procedures34,35 were modified to obtain the regioisomeric mixture of
1(1,6/1,7). Until 2004,28 the existence of the isomeric mixture of 1 was not known, since the
two distinct regioisomers could only be analyzed with NMR spectroscopy at a field strength
of 400 MHz or higher. However, this could be expected due to the known intermolecular
aggregation of PDIs as a result of π –π stacking. It has been shown that bulky groups
impede intermolecular aggregation.36 Consistent with our previous report,34 to prevent ag-
gregation and achieve the facile separation of the two isomers, we elected to append bulkier
moieties to 1(1,6/1,7). We chose the triphenylsilylacetylene (TPSA) group for the bulky
bay substituent and the 1-pentynyl group for the small group. The PDI mixture 1(1,6/1,7)
reacted with TPSA under standard Sonogashira cross-coupling conditions (Scheme 1). As
expected, the subsequent regioisomers showed a definite difference in their Rf values in a
toluene:hexane (19:1) solvent system and thus were easily separated by column chromatog-
raphy. The first bright orange band contained 2(1,6) in 4% yield, and the last bright orange
band contained 3(1,7) in 33% yield. However, there was minimal separation of the two
bands, and 42% of the material loaded on the column was recovered as the regioisomeric
mixture. With TPSA groups at the bay region, all conditions used to convert the diimides
back to dianhydrides for subsequent imide functionalization were unsuccessful due to the

Scheme 1
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 741

Scheme 2
Downloaded by [Tulane University] at 06:39 08 January 2015

saponification conditions required for conversion (KOH, isopropanol, 83 ◦ C, 2 h). Under


basic conditions, the TPSA groups were desilylated affording a totally insoluble material.
Thus, for comparison the effects of tritylacetylene groups34 were examined.
Bulky groups affect the solid state π –π stacking and consequent electron mobility as
Anthony and coworkers elegantly showed with substituted pentacenes.36 Thus, we elected
to synthesize regioisomerically pure PDIs with long side chains. Therefore, 1(1,6/1,7) was
allowed to react with 1-pentyne by standard Sonogashira coupling (Scheme 1) and yielded
the regioisomeric mixtures of 4/5(1,6/1,7). The resulting isomers of 4/5(1,6/1,7) showed
no Rf difference in a variety of solvent combinations and could not be separated by column
chromatography. However, the pentyne groups at the bay region of 4/5(1,6/1,7) could be
reduced via the palladium-catalyzed hydrogenation to afford the chromatographically sep-
arable regioisomers of 6(1,6) and 7(1,7) (Scheme 2). Unfortunately, due to the very limited
solubility of 6(1,6) and 7(1,7), it was not possible to convert them to the corresponding
PTCDs. For later discussion and the sake of comparison, Scheme 3 illustrates a series of
1,6- and 1,7-regioisomers of PDIs and PTCDs most of which have not been subjected to
detailed analysis.34

Scheme 3
742 NISHA V. HANDA ET AL.

Figure 1 Labeled protons of 1,7- and 1,6-PDIs.


Downloaded by [Tulane University] at 06:39 08 January 2015

The 1H-NMR spectra of 1,6- and 1,7-regioisomers of all PDIs appended with alkyl
groups (6, 7, 834, and 934) revealed considerable differences in chemical shifts as well
as the splitting pattern of the aromatic protons. The 1,7-regioisomers, 7 and 9 exhibited
significant upfield and downfield shifts of ∼0.07 ppm for Hc and Ha (for labeled protons,
see Figure 1), respectively, as compared to the shifts in the corresponding 1,6-regioisomers,
6 and 8. In contrast, the 1H-NMR spectra of tritylacetylene derivatives did not show a
significant difference except for a slight upfield shift of 0.04 ppm for Hc in 10(1,6) as
compared to that in 11(1,7) (see SI for spectrum). An evaluation of the chemical shifts for
Hb and Hc in 3(1,7) and 10(1,6) suggests that Hb probably lies in the deshielding region
of the triple bond. This is supported by the observation in the n-hexyl derivative 12(1,7),
which is void of the triphenylmethyl group, that only Hb is positioned close to the triple
bond and is therefore highly deshielded. In contrast, Hc is much more shielded in 3(1,7)
and 10(1,6), being situated outside the deshielding area of the triple bond. The deshielding
effect of triple bonds on nearby protons in the rigid system is documented.37
To determine the HOMO and LUMO levels of all PDIs and PTCDs synthesized, their
electrochemical properties were investigated by cyclic voltammetry. The first and second
half wave reduction potentials, the position of HOMO/LUMO levels, and the optical band
gap (Eg ) values are depicted in Table 1. All derivatives underwent two successive reversible
reductions corresponding to the radical anion and dianion.38 No reversible oxidation was
observed for these derivatives within the scanning range, which the conditions allowed, as
was expected according to previously published reports.1,2 Only PDI or PTCD cores with
electron donating bay substituents have been shown to undergo reversible oxidation.39,40
In accordance with the literature,29,32 the reductive capability of both regioisomers of a
series of PDIs and PTCDs did not exhibit significant differences. However, the first and
second reduction potential of 2(1,6) was found to be −10 mV and −20 mV lower than the
corresponding 3(1,7), respectively. Similar to this, the reduction potentials for 14(1,6) were
−20 mV lower than the 15(1,7).
LUMO and HOMO energy levels of all PTCDs were found to be in the range of −4.2
to −3.8 eV and −6.4 to −6.2 eV, respectively, demonstrating that these are potentially
suitable candidates for electron transport materials in electronic devices.
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 743

Table 1 Electrochemical properties of 1,7- and 1,6-PDIs and PTCDs

Entry E−11/2 [V vs. Fc/Fc+]a E2−1/2 [V vs. Fc/Fc+]a Eg (eV)b ELUMO (eV)c EHOMO (eV)d

2(1,6) −0.91 −1.24 2.3 −4.1 −6.4


3(1,7) −0.90 −1.22 2.3 −4.1 −6.4
4/5(1,6/1,7) −1.00 −1.33 2.3 −4.0 −6.3
6(1,6) −1.18 −1.51 2.4 −3.8 −6.2
7(1,7) −1.16 −1.50 2.4 −3.8 −6.2
8(1,6) −1.16 −1.49 2.4 −3.8 −6.2
9(1,7) −1.16 −1.50 2.4 −3.8 −6.2
10(1,6) −0.96 −1.30 2.3 −4.0 −6.3
11(1,7) −0.96 −1.31 2.2 −4.0 −6.2
12(1,7) −1.01 −1.33 2.3 −4.0 −6.3
13 −1.1 −1.42 2.4 −3.9 −6.3
14(1,6) −0.97 −1.26 2.4 −4.0 −6.4
15(1,7) −0.95 −1.24 2.4 −4.0 −6.4
16(1,6) −0.79 −1.10 2.2 −4.2 −6.4
Downloaded by [Tulane University] at 06:39 08 January 2015

17(1,7) −0.79 −1.11 2.2 −4.2 −6.4

a 0.1 M TBAPF in THF. b E = (hc)/λ


max = optical band gap. ELUMO = 4.4 eV − Ered1 . EHOMO = ELUMO
c d
4 g
− Eg .

The effect of bay substitution on electron affinity is discussed relative to unsubstituted


PDI 13. It is widely accepted that electronics of bay-substituents play a major and decisive
role in redox processes. Electron-withdrawing groups enhance the electron affinity of
resulting PDI derivatives whereas electron-releasing groups diminishes it.1 However, in
our system, neutral groups such as pentyne and hexyne at the bay region of PDIs in
4/5(1,6/1,7) and 12(1,7) enhanced the reductive capability of resulting PDI derivatives by
∼ 100 mV in contrast to the corresponding unsubstituted analogue 13. This effect was
probably due to an extended conjugation imparted by the alkyne moieties directly attached
at the bay region.4 Furthermore, upon substitution of the bay region of PDI with neutral,
bulky, and conjugated groups including triphenylsilylacetylene and tritylpropynyl in 3(1,7)
and 11(1,7), the reductive capability is significantly enhanced by ∼140 mV–200 mV. This
suggests that, along with the electronics, an extended conjugation and steric effects imparted
by the bay groups also exhibit a major role in redox processes. In contrast, the pentyl and
hexyl groups in 7(1,7) and 9(1,7), despite being neutral and nonconjugated, were found to
slightly lower the reductive capability of the resulting PDI derivative by ∼60 mV. Therefore,
although it is difficult to ascertain the extent of involvement of both electronic and steric
factors in redox processes, the role of steric effects cannot be completely ruled out.
The PTCD derivatives appended with hexyl and tritylacetylene groups were found to
reduce significantly easier by ∼200 mV in comparison to the corresponding PDI derivatives.
This observation could reasonably be explained in terms of higher electronegativity of the
oxygen in PTCDs in contrast to the nitrogen in PDI derivatives.

STRUCTURAL PROPERTIES AND TWIST ANGLES


To compare the solid-state structures of 1,6- and 1,7-regioisomers of bay appended
PDIs, X-ray crystallography studies were performed. Single red plate and orange block
shaped crystals of 10(1,6) and 11(1,7) were grown by slow vapor diffusion of pentane
into a CH2 Cl2 solution of each compound. PDI derivative 11(1,7), consisting of bulky
744 NISHA V. HANDA ET AL.

Figure 2 Molecular structure of 11(1,7) in the crystals with numbering of atoms (left) and side view along N–N
axis. Thermal ellipsoids are set at 50% probability.

tritylpropynyl groups at the bay region, crystallized in the monoclinic space group P21 /c. To
Downloaded by [Tulane University] at 06:39 08 January 2015

our surprise, regardless of the bulky triphenyl rings, the central PDI core assumed an almost
planar conformation, equivalent to an unsubstituted PDI.41 The torsion angle associated with
the bay C atoms C(1)-C(7)-C(9)-C(6) and C(6)-C(9)-C(7)-C(1) (see Figure 2 for carbon
numbering) was found to be similar, with a value of 1.3(3)◦ , which is slight less than the
torsion angle of the bay carbon atoms in the unsubstituted PDI (1.8◦ ).41 Some literature
precedents indicate a nearly planar π -conformation of bay appended PDIs.3,42–44 However,
there is only one very recent example published which includes a distorted core of a PDI
unit bearing bulky diphenylphenoxy groups.45
A tentative explanation for the flat PDI core might be that the elongated tritylacetylene
groups allow the large trityl units to be accommodated at sufficient distance from C-H6
to minimize any compression effect on the core, thereby reducing twisting of the internal
six-membered rings. The staggered arrangement of the large tritylacetylene groups across
the core likely contributes to alleviate strain.
The corresponding 10(1,6), crystallized in the monoclinic space group C2/c. A careful
examination of the 1,6-regioisomer 10 (Figure 3) revealed reduced angles of 113.1(5)◦ for
the C(2)-C(1)-C(35) and 116.6(4)◦ for C(5)-C(6)-C(56) arrangements. This compression
effect resulted in enlarged angles for C(1)-C(13)-C(18) and C(6)-C(15)-C(16) of 123.8(5)◦
and 123.4(4)◦ , respectively. Interestingly, the torsion angles associated with the bay area
C atoms C(1)-C(13)-C(18)-C(12) and C(6)-C(15)-C(16)-C(7) were found to be 12.8(7)◦
and 19.4(7)◦ , respectively, implies that considerable unsymmetrical twisting exists of the
central core of six- membered rings. This is reasonable if the large tritylacetylene groups
turn away from C-H7 and C-H12 bonds to minimize repulsive forces, resulting in part of
the core being out of planarity with the remaining central unit, which appears directly in
the crystal lattice.
Some information exists on the crystal structures of compounds containing triaryl
units.46,47 In our systems, the triphenyl rings in 10(1,6) and 11(1,7) assumed a propeller-
like configuration, and consequently, adopted a six-fold phenyl embrace (6-PE) interaction.
This probably occurred as a result of complementary parallel interaction involving adjacent
Ph3 X groups with a stabilization energy of about 50–80 kJ/mol.48 However, the reduced
twisting in 11(1,7) and 10(1,6) might also be attributed to specific packing effects imparted
by the strong supramolecular 6-PE interaction.
The orange block-shaped crystals of 9(1,7) were grown by vapor diffusion of MeOH
into a THF solution. Such data allow a comparison of the effect of nonbulky hexyl group on
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 745
Downloaded by [Tulane University] at 06:39 08 January 2015

Figure 3 Molecular structure of 10(1,6) in the crystals with numbering of atoms. Thermal ellipsoids are set at
50% probability.

Figure 4 Molecular structure of 9(1,7) in the crystals with numbering of atoms. Thermal ellipsoids are set at
50% probability.
746 NISHA V. HANDA ET AL.

Figure 5 The molecular structure of 17(1,7) in the crystals with numbering of atoms. Thermal ellipsoids are set
at 50% probability.
Downloaded by [Tulane University] at 06:39 08 January 2015

the twisting and resulting packing in the solid state. The solid crystallized in an orthorhombic
space group Pbca with a torsion angle of 21.8(6)◦ and 24.2(6)◦ involving the bay C atoms
C(6)-C(15)-C(16)-C(7) and C(1)-C(13)-C(18)-C(12) due to the unsymmetrical twisting of
central six membered ring (Figure 4) which is comparable to that (26.4◦ and 29.6◦ ) of
a core-perfluoroalkylated PDI.49 The significant twisting of the PDI core was somewhat
expected due to the steric hindrance caused by the dangling hexyl chains at the bay region.
We were not able to obtain useful single crystals of 8(1,6) even after many attempts.
Red, needle-shaped crystals of 17(1,7) were collected and analyzed. The triphenyl
rings in the bisanhydride 17(1,7) assumed a propeller-like configuration in the solid state
as found in 11(1,7). The triclinic space group P1 had a torsion angle of −9.2 (2) for C(1)-
C(7)-C9)-C6) and is similar to that found in 11(1,7). Minor twisting exists in bisanhydride
11(1,7) (Figure 5).

Optical Properties
The solution state absorption spectrum of 1,6- and 1,7-regioisomers of triphenylsilyl
and pentyl appended PDIs were found to be in general agreement with our previously
reported PDIs.34 Both 1,6- and 1,7-regioisomers of 2 and 3 were found to possess specific
characteristic absorbances at ∼300 nm and 350 nm, respectively (Figure 6). See also data
in Table 2.

Table 2 Optical properties of 1,7- and 1,6-regioisomers of PDIs

Entry λmax (nm)a ε (M−1 cm−1)b λmax (nm)c λemi (nm)d λemi (nm)e

2(1,6) 547 60972 555 565 645


3(1,7) 550 55745 558 570 645
4/5(1,6/1,7) 548 48699 573 568 648
6(1,6) 520 54441 527 541 610
7 1,7) 520 59226 528 543 595

a Absorption maxima in CH Cl (10−5 M). b Determined at λ c


2 2 max in the solution state. Absorption maxima in
the solid state (10−3 M). d Emission maxima in CH2 Cl2 . e Emission maxima in the solid state.
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 747
Downloaded by [Tulane University] at 06:39 08 January 2015

Figure 6 Steady-state absorption spectrum in solution state (1 × 10−5 M, CH2 Cl2, 25 ◦ C). Red: R = TPSA,
Black: R = pentyl. 1,7-regioisomer = solid line; 1,6-regioisomer = dotted line.

After casting all PDIs and PTCDs from solution to solid films, the full width at
half-maximum (FWHM) in the absorption spectrum was found to increase and to become
much broader (see SI for spectrum), which suggested the presence of an improved stacked
structure in the solid state than in solution.50 Additionally, the solid-state absorption spectra
of alkynyl appended PDIs and PTCDs shifted bathochromically by ∼8–25 nm in compari-
son to the solution state (Tables 2 and SI). The characteristic major absorption of the 1,7-
and 1,6-regioisomers at ∼300 nm and 350 nm, respectively, was also found in a majority
of cases to undergo a bathochromic shift (5–25) nm), relative to the solution state (SI).
Both 14/15(1,6/1,7) showed a significant bathochromic shift of 25–31 nm in comparison
to the solution state spectra. The bathochromic shift of all absorptions in the solid state
can probably be attributed to increased aggregation and to the different intermolecular
π -interactions in the solid state.51
The emission spectrum of all PDIs and PTCDs in the solid state exhibited a significant
bathochromic shift of 60–150 nm with a larger Stokes shift of ∼80–130 nm, relative to
the solution state spectrum (see Tables 2 and SI). The shift was likely due to molecular
stacking and aggregation in the solid state. The shape of the fluorescence spectra was also
found to be different. There was no distinguishable shoulder peak corresponding to S2 –S0
transition as in the solution, thereby suggesting the presence of pronounced intermolecular
interactions in the solid state.52

CONCLUSIONS
A small series of 1,6- and 1,7-regioisomers of PDIs incorporated with different sized
groups were synthesized, separated by conventional column chromatography, and correlated
with each other and related systems. The tritylacetylene appended PDI, despite the presence
of the bulky group, assumed an almost planar π -configuration of the central naphthalene
moiety with a twist angle of only 1.5(5)◦ . To date, there is only one report available in the
748 NISHA V. HANDA ET AL.

literature for an undistorted PDI core consisting of bulky groups.53 Interestingly, in our work
the PDI incorporated with the slender hexyl chain exhibited an unsymmetrical twisting of
21.8(6)◦ and 24.2(6)◦ . The optical data were also found to be in good correlation with
the twist angle. The PDI containing a hexyl chain undergoes a significant hypsochromic
shift due to the twisting. We have shown that, along with the electronics, sterics of the bay
substituents are also major contributing factors in determining the reductive capability of
a PDI. In the solid state, a bathochromic shift of all absorptions with all PDIs and PTCDs
was observed which could be due to enhanced aggregation. Moreover, despite a significant
bathochromic shift in λemi , fluorescence quenching was observed in the solid state. The
difference in electron mobility of 1,6- and 1,7-regioisomers of different substituted PTCDs
will be published in due course.

EXPERIMENTAL
Downloaded by [Tulane University] at 06:39 08 January 2015

All solvents and reagents were purchased from a commercial supplier and used as
received unless otherwise indicated. THF and CH2 Cl2 were dried immediately prior to
use via a SP-105 solvent purification system by LC Technology Solutions. Silica TLC
plates (UV 254 indicator) and silica gel (230–400 mesh) for column chromatography were
purchased from Sorbent technologies. All NMR solvents were purchased from Cambridge
Isotopes Laboratories and used as received. All NMR spectra (1H-NMR and 13C-NMR)
were recorded on a Varian Unity INOVA (400 MHz) spectrometer. Chemical shifts (δ) are
expressed in ppm relative to the residual chloroform (1H: 7.26 ppm, 13C: 77.16 ppm)
as a reference. Melting points were determined with Stuart SMP10 instrument using
Kimble Kimex 51 capillaries (1.5–1.8 × 90 mm) and were uncorrected. Fourier trans-
form infrared (IR) measurements were performed on Varian 800 FT-IR spectrometer. A
Cary 5000 Bio UV–VIS–NIR spectrophotometer was used to record the UV–VIS spec-
tra. A Cary eclipse fluorescence spectrophotometer was used to obtain emission spectra.
The X-ray diffraction data were collected with a Bruker APEX CCD area detector us-
ing graphite-monochromated Mo Kα radiation (λ = 0.71073 Å).54 The voltamograms
were determined with cyclic voltammetry CV-50W instrument using three-electrode single
compartment cell consisting of Glassy carbon as the working electrode, Pt as the counter
electrode, and silver–silver chloride as the reference electrode under a continuous flow
of helium gas. All measurements were performed at 100 mV/s, and a 0.1 M solution
of tetrabutyl-ammonium tetrafluoroborate in THF was used as the supporting electrolyte.
High-resolution mass spectra were analyzed on hybrid LTQ-Orbitrap XL mass spectrome-
ter. N,N’-Bis(n-ethylpropyl)perylene-3,4:9,10-tetracarboxylic diimide (13), and a mixture
of 1,6- and 1,7-dibromo-N,N’-bis(ethylpropyl)perylene-3,4:9,10-tetracarboxylic diimides
(1) were synthesized according to a literature procedures.35 Compounds 9--12,34 14--17,34
and 1328 have been recorded, but such have not been examined for potential optical or elec-
trical properties. Supplementary Materials (Figures S1–S25) contain spectra and analysis
of the new compounds.

GENERAL SONOGASHIRA COUPLING PROCEDURE A


To a clean, dry, three-necked, round-bottomed flask equipped with a septum and a stir
bar was added the aromatic bromide (1.0 equiv.) dissolved in Et3 N (0.02 M). After purging
with argon, Pd(PPh3 )4 (5.0 mol %), CuI (2.5 mol %), and the appropriate alkyne (3.0 equiv)
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 749

were added. The solution was refluxed at 90 ◦ C under Ar for 24 h. The reaction mixture
was filtered through celite, and the crude product was concentrated in vacuo. The crude
material was washed with methanol and purified by column chromatography on silica gel
to afford the desired product.

GENERAL REDUCTION PROCEDURE B


To a stirred solution of alkyne (1.0 equiv) in a mixture of CH2 Cl2 :MeOH (2:1,
0.01 M) under an argon-filled balloon was added 10% Pd/C (10%–20% by weight) and
neat triethylsilane (10 equiv). Completion of the reaction was monitored by a color change
from red to green. After stirring (1 h), the mixture was filtered through a short pad of
celite, and the solvent was removed to give the desired product whose molecular weight
was confirmed by HR mass spectrometry.
Downloaded by [Tulane University] at 06:39 08 January 2015

1,6-Bis(Triphenylsilylacetylene)-N,N’-bis(ethylpropyl)perylene-3,4:9,10-
tetracarboxylic Acid Bisimide 2(1,6) and 1,7-Bis(Triphenylsilylacetylene)-
N,N’-bis-(ethyl-propyl)perylene-3,4:9,10-tetracarboxylic Acid Bisimide 3(1,7)
A mixture of compounds 134,35(1,6/1,7) (0.44 g, 0.65 mmol) were treated with
Pd(PPh3 )4 (37.3 mg, 0.03 mmol), CuI (3.2 mg, 0.02 mmol), and triphenylsilylacetylene
(0.55 g, 1.94 mmol) in Et3 N (30 mL) following general procedure A. The crude product
was purified by column chromatography (toluene/hexane (19:1). Isolation of the first band
afforded pure 2 (1,6) (10 mg, 4.2%) as a red solid. mp > 300 ◦ C; 1H NMR (400 MHz,
CDCl3 ): δ 10.02 (d, J = 8 Hz, 2H), 8.93 (s, 2H), 8.25(d, J = 8 Hz, 2H), 7.77-7.74 (m,
12H), 7.55-7.45 (m, 18H), 5.08-5.04 (m, 2H), 2.31-2.23 (m, 4H), 1.99-1.89 (m, 4H), 0.95-
0.88 (m, 12H); 13C-NMR (100 MHz, CDCl3 ): δ 138.9, 135.9, 135.8, 133.0, 132.4, 130.6,
128.4, 128.1, 128.0, 127.9, 127.1, 119.9, 109.4, 100.2, 58.1, 57.7, 25.1, 11.4; IR(neat):
2957, 2931, 2873, 2142, 1697, 1654; HRMS (ESI) for C74 H58 N2 O4 Si2 [M++H]: Calcd
for 1095.4013. Found: 1095.4005. Isolation of the second band afforded 3 (1,7) (0.24 g,
33%) as red solid. mp > 300 ◦ C; 1H NMR (400 MHz, CDCl3 ): δ 10.06 (d, J = 8.4 Hz,
2H), 8.93 (s, 2H), 8.26 (d, J = 8.4 Hz, 2H), 7.78 (d, J = 6.8 Hz, 12H), 7.54-7.47 (m,
18H), 5.07 (qt, J = 4 Hz, 2H), 2.32–2.24 (m, 4H), 1.95 (qt, J = 6.4Hz, 4H), 0.93(t, J =
7.6Hz, 12H); 13C-NMR (100 MHz, CDCl3 ): δ 164.0, 138.1, 136.1, 135.8, 135.6, 135.0,
133.8, 132.4, 131.5, 130.6, 128.7, 128.4, 128.2, 128.2, 127.5, 123.5, 122.5, 119.4, 109.4,
100.5, 57.9, 25.1, 11.4; IR(neat): 2957, 2933, 2873, 2144, 1699, 1655; HRMS (ESI) for
C74 H58 N2 O4 Si2 [M++H]: Calcd for 1095.4013. Found: 1095.4000.

1,6-Dipentyne-N,N’-bis(ethylpropyl)perylene-3,4:9,10-tetracarboxylic
Acid Bisimide and 1,7-Dipentyne-N,N’-bis(ethylpropyl)perylene-3,4:9,10-
tetracarboxylic Acid Bisimide 4/5(1,6/1,7)
A mixture of compound 1 (1,6/1,7) (1.0 g, 1.45 mmol), Pd(PPh3 )4 (84 mg, 0.07 mmol),
CuI (7.0 mg, 0.04 mmol), 1-pentyne (0.36 mL, 3.62 mmol), and triethylamine (60 mL,
0.02 M) was allowed to react according to the general procedure A. The crude product
was purified by column chromatography (CH2 Cl2 :hexane, 3:2) to afford a mixture of 4/5
(1,6/1,7) (0.7 g, 73%). mp > 300 ◦ C; 1H-NMR (400 MHz, CDCl3 ): δ 10.13 (d, J = 8.4 Hz,
2H), 8.76 (s, 2H), 8.63 (d, J = 8.4 Hz, 2H), 5.10-5.03 (m, 2H), 2.63 (t, J = 7.0 Hz, 4H),
2.33-2.21 (m, 4H),1.97-1.90 (m, 4H), 1.84-1.75 (m, 4H), 1.17 (t, J = 7.4 Hz, 6H), 0.92 (t,
750 NISHA V. HANDA ET AL.

J = 7.4 Hz, 12H); 13C-NMR (100 MHz, CDCl3 ): δ 165, 138.3, 134.2, 133.8, 130.5, 127.7,
127.0, 126.6, 121.7, 121.0, 101.3, 82.6, 57.8, 25.2, 25.1, 22.4, 22.0, 14.0, 11.5; IR(neat):
2960, 2929, 2870, 2209, 1695, 1650; HRMS (ESI) for C44 H42 N2 O4 [M++H]: Calcd for
663.3223. Found: 663.3206. [M++2 H]: Calcd for 664.3301. Found: 664.3234.

1,6-Dipentyl-N,N’-bis(ethylpropyl)perylene-3,4:9,10-tetracarboxylic
Acid Bisimide 6(1,6)- and 1,7-Dipentyl-N,N’- bis (ethyl propyl) perylene-
3,4:9,10-tetracarboxylic Acid Bisimide 7(1,7)
A mixture of 4/5(1,6/1,7) (0.67 g, 1 mmol) and 10% Pd/C (134 mg) were treated with
triethylsilane (1.6 mL, 10 mmol) following general procedure B. The product was purified
by column chromatography using (CH2 Cl2 :hexane (1:1)). Isolation of the first compound
eluted afforded pure 6 (1,6) (70 mg, 10%) as a bright red solid, mp = 180–181 ◦ C; 1H-NMR
(400 MHz, CDCl3 ): δ 8.70 (d, J = 7.60 Hz, 2H), 8.63 (s, 2H), 8.26 (d, J = 8.0 Hz, 2H),
5.13-5.06 (m, 2H), 3.29-3.25 (m, 4H), 2.34-2.22 (m, 4H), 2.05-1.90 (m, 8H), 1.54-1.42
Downloaded by [Tulane University] at 06:39 08 January 2015

(m, 8H), 0.98-0.89 (m, 18H); 13C-NMR (100 MHz, CDCl3 ): δ 164.3, 142.2, 134.6, 133.7,
132.9, 131.0, 130.1, 128.9, 128.8, 128.5, 128.4, 127.5, 125.4, 122.3, 57.7, 57.6, 36.3, 32.1,
31.2, 25.2, 25.1, 22.5, 14.2, 11.5; IR(neat): 2954, 2924, 2855, 1692, 1649; HRMS (ESI)
for C44 H50 N2 O4 [M+-H]: Calcd for 669.3692. Found: 669.3671; [M+] Calcd for 670.3770.
Found: 670.3720. Isolation of the second band afforded 7 (1,7) (110 mg, 16%) as bright red
solid, mp = 192–193 ◦ C; 1H-NMR (400 MHz, CDCl3 ): δ 8.69 (s, 2H), 8.63 (d, J = 8.0 Hz,
2H), 8.26 (d, J = 8.0 Hz, 2H), 5.13-5.05 (m, 2H), 3.34-3.30 (m, 4H), 2.35-2.22 (m, 4H),
2.06-1.89 (m, 8H), 1.53-1.41 (m, 8H), 0.98-0.91 (m,18H); 13C NMR (100 MHz, CDCl3 ): δ
164.3, 141.4, 134.5, 133.4, 129.5, 128.7, 128.2, 127.0, 122.3, 57.7, 36.0, 32.1, 31.0, 25.2,
22.5, 14.2, 11.5; IR(neat): 2954, 2929, 2869, 1693, 1650; HRMS (ESI) for C44 H50 N2 O4
[M+-H]: Calcd: 669.3692. Found: 669.3671.

SUPPLEMENTAL MATERIAL
Supplementary data of this article can be accessed on the publisher’s website,
www.tandfonline.com/gpss

ACKNOWLEDGMENT
We acknowledge the College of Arts and Sciences and the Department of Chemistry
at Oklahoma State University for support. We thank Dr. Sachin Handa for help in crystal
growth and Dr. Steve Hartson for assistance in HRMS. Crystallographic data for 9, 10,
11, and 17 are deposited with the Cambridge depository as CCDC947158, CCDC947159,
CCDC940953, and CCDC 940954, respectively.

FUNDING
We acknowledge funding for the Varian Inova 400 MHz NMR spectrometer in the
Oklahoma Statewide Shared NMR facility by the National Science Foundation (BIR-
9512269), the Oklahoma State Regents for Higher Education, the W. M. Keck Foundation,
and Conoco, Inc. We acknowledge the National Science Foundation (CHE-0130835) and
the University of Oklahoma for funds to purchase the X-ray unit and computers.
1,6- AND 1,7-REGIOISOMERS OF PERYLENE TETRACARBOXYLIC DIANHYDRIDE 751

REFERENCES
1. Würthner, F. Chem. Commun. 2004, (14):1564-1579.
2. Huang, C.; Barlow, S.; Marder, S. R. J. Org. Chem. 2011, 76, 2386-2407.
3. Jones, B. A.; Ahrens, M. J.; Yoon, M.-H.; Facchetti, A.; Marks, T. J.; Wasielewski, M. R. Angew.
Chem. Int. Ed. 2004, 43, 6363-6366.
4. Dimitrakopoulos, C. D.; Malenfant, P. R. L. Adv. Mater. 2002, 14, 99-117.
5. Cespedes-Guirao, F. J.; Ropero, A. B.; Font-Sanchis, E.; Nadal, A.; Fernandez-Lazaro, F.; Sastre-
Santos, A. Chem. Commun. 2011, 47, 8307-8309.
6. Peneva, K.; Mihov, G.; Herrmann, A.; Zarrabi, N.; Boersch, M.; Duncan, T. M.; Muellen, K. J.
Am. Chem. Soc. 2008, 130, 5398-5399.
7. Peneva, K.; Mihov, G.; Nolde, F.; Rocha, S.; Hotta, J.-i.; Braeckmans, K.; Hofkens, J.; Uji-i, H.;
Herrmann, A.; Müllen, K. Angew. Chem. Int. Ed. 2008, 47, 3372-3375
8. Yukruk, F.; Dogan, A. L.; Canpinar, H.; Guc, D.; Akkaya, E. U. Org. Lett. 2005, 7, 2885-2887.
9. Choi, M. J.; Smoother, T.; Martin, A. A.; McDonagh, A. M.; Maynard, P. J.; Lennard, C.; Roux,
C. Forensic Sci. Int. 2007, 173, 154-160.
Downloaded by [Tulane University] at 06:39 08 January 2015

10. Szelke, H.; Schubel, S.; Harenberg, J.; Kramer, R. Chem. Commun. 2010, 46, 1667-1669.
11. Szelke, H.; Schübel, S.; Harenberg, J.; Krämer, R. Bioorg. Med. Chem. Lett. 2010, 20, 1445-1447.
12. Rodrı́guez-Morgade, M. S.; Torres, T.; Atienza-Castellanos, C.; Guldi, D. M. J. Am. Chem. Soc.
2006, 128, 15145-15154
13. Würthner, F.; Kaiser, T. E.; Saha-Möller, C. R. Angew. Chem. Int. Ed. 2011, 50, 3376-3410.
Safot-Sempere, M. M.; Fernandez, G.; Würthner, F. Chem. Rev. 2011, 111, 5784–5814.
14. Wang, B.; Yu, C. Angew. Chem. Int. Ed. 2010, 49, 1485-1488.
15. Chen, Z.; Debije, M. G.; Debaerdemaeker, T.; Osswald, P.; Würthner, F. Chem. Phys. Chem.
2004, 5, 137-140.
16. Tang, C. W. Appl. Phys. Lett. 1986, 48, 183-185.
17. Schmidt-Mende, L.; Fechtenkotter, A.; Mullen, K.; Moons, E.; Friend, R. H.; MacKenzie, J. D.
Science 2001, 293, 1119-1122.
18. Angadi, M. A.; Gosztola, D.; Wasielewski, M. R. Mater. Sci. Eng., B 1999, B63, 191-194.
19. Queste, M.; Cadiou, C.; Pagoaga, B.; Giraudet, L.; Hoffmann, N. New J. Chem. 2010, 34,
2537-2545.
20. Chao, C.-C.; Leung, M.-k.; Su, Y. O.; Chiu, K.-Y.; Lin, T.-H.; Shieh, S.-J.; Lin, S.-C. J. Org.
Chem. 2005, 70, 4323-4331.
21. Zhao, Y.; Wasielewski, M. R. Tetrahedron Lett. 1999, 40, 7047-7050.
22. Nakazono, S.; Easwaramoorthi, S.; Kim, D.; Shinokubo, H.; Osuka, A. Org. Lett. 2009, 11,
5426-5429.
23. Nakazono, S.; Imazaki, Y.; Yoo, H.; Yang, J.; Sasamori, T.; Tokitoh, N.; Cédric, T.; Kageyama,
H.; Kim, D.; Shinokubo, H.; Osuka, A. Chem. Eur. J. 2009, 15, 7530-7533.
24. Langhals, H.; Demmig, S.; Huber, H. Spectrochim. Acta, Part A 1988, 44A, 1189-1193.
25. Rohr, U.; Kohl, C.; Mullen, K.; van, d. C. A.; Warman, J. J. Mater. Chem. 2001, 11, 1789-1799.
26. Wescott, L. D.; Mattern, D. L. J. Org. Chem. 2003, 68, 10058-10066.
27. Würthner, F.; Thalacker, C.; Diele, S.; Tschierske, C. Chem. Eur. J. 2001, 7, 2245-2253.
28. (a) Würthner, F.; Stepanenko, V.; Chen, Z.; Saha-Moeller, C. R.; Kocher, N.; Stalke, D. J. Org.
Chem. 2004, 69, 7933-7939. (b) Bohm, A.; Arms, H.; Henning, G.; Blaschka, P. (BASF AG)
German Patent DE19547209 A1, 1997; Chem. Abstr. 1997, 127, 96569g.
29. Keerthi, A.; Valiyaveettil, S. J. Phys. Chem. B 2012, 116, 4603-4614.
30. Zhao, Q.; Zhang, S.; Liu, Y.; Mei, J.; Chen, S.; Lu, P.; Qin, A.; Ma, Y.; Sun, J. Z.; Tang, B. Z. J.
Mater. Chem. 2012, 22, 7387-7394.
31. Dey, S.; Efimov, A.; Lemmetyinen, H. Eur. J. Org. Chem. 2012, 2012, 2367-2374.
32. Dubey, R. K.; Efimov, A.; Lemmetyinen, H. Chem. Mater. 2011, 23, 778-788.
33. Jimenez, A. J.; Spaenig, F.; Rodriguez-Morgade, M. S.; Ohkubo, K.; Fukuzumi, S.; Guldi, D.
M.; Torres, T. Org. Lett. 2007, 9, 2481-2484.
752 NISHA V. HANDA ET AL.

34. Handa, N. V.; Mendoza, K. D.; Shirtcliff, L. D. Org. Lett. 2011, 13, 4724-4727.
35. (a) Rajasingh, P.; Cohen, R.; Shirman, E.; Shimon, L. J. W.; Rybtchinski, B. J. Org. Chem. 2007,
72, 5973-5979. (b) Demmig, S.; Langhals, H. Chem. Ber. 1988, 121, 225-230.
36. Anthony, J. E.; Eaton, D. L.; Parkin, S. R. Organic Lett. 2002, 4, 15-18.
37. Abraham, R. J.; Reid, M. J. Chem. Soc. 2001, 7 1195-1204.
38. Lee, S. K.; Zu, Y.; Herrmann, A.; Geerts, Y.; Müllen, K.; Bard, A. J. J. Am. Chem. Soc. 1999,
121, 3513-3520.
39. Zhao, Y.; Wasielewski, M. R. Tetrahedron Lett. 1999, 40, 7047-7050.
40. Lukas, A. S.; Zhao, Y.; Miller, S. E.; Wasielewski, M. R. J. Phys. Chem. B 2002, 106, 1299-1306.
41. Headicke, E.; Graser, F. Acta Crystallogr. Sec. 1986, C42, 189-195.
42. Schmidt, R.; Ling, M. M.; Oh, J. H.; Winkler, M.; Könemann, M.; Bao, Z.; Würthner, F. Adv.
Mater. 2007, 19, 3692-3695.
43. Delgado, M. C. R.; Kim, E.-G.; Filho, D. t. A. d. S.; Bredas, J.-L. J. Am. Chem. Soc. 2010, 132,
3375-3387.
44. Weissman, H.; Shirman, E.; Ben-Moshe, T.; Cohen, R.; Leitus, G.; Shimon, L. J. W.; Rybtchinski,
B. Inorg. Chem. 2007, 46, 4790-4792.
Downloaded by [Tulane University] at 06:39 08 January 2015

45. Lin, M.-J.; Jimenez, A. J.; Burschka, C.; Würthner, F. Chem. Commun. 2012, 48, 12050-12052.
46. Karlen, S. D.; Khan, S. I.; Garcia-Garibay, M. A. Crystal Growth Des. 2005, 5, 53-55.
47. Scudder, M.; Dance, I. J. Chem. Soc. 1998, (3):329-344.
48. Janssen, F. F. B. J.; de Gelder, R.; Rowan, A. E. Crystal Growth Des. 2011, 11, 4326-4333.
49. Li, Y.; Tan, L.; Wang, Z.; Qian, H.; Shi, Y.; Hu, W. Org. Lett. 2008, 10, 529-532.
50. Shin, W. S.; Jeong, H.-H.; Kim, M.-K.; Jin, S.-H.; Kim, M.-R.; Lee, J.-K.; Lee, J. W.; Gal, Y.-S.
J. Mater. Chem. 2006, 16, 384-390.
51. Amiralaei, S.; Uzun, D.; Icil, H. Photochem. Photobiol. Sci. 2008, 7, 936-947.
52. Ma, Y.-S.; Wang, C.-H.; Zhao, Y.-J.; Yu, Y.; Han, C.-X.; Qiu, X.-J.; Shi, Z. Supramol. Chem.
2007, 19, 141-149.
53. Ramanan, C.; Smeigh, A. L.; Anthony, J. E.; Marks, T. J.; Wasielewski, M. R. J. Am. Chem. Soc.
2012, 134, 386-397.
54. (a) Data Collection: Bruker. SMART Bruker AXS Inc: Madison, Wisconsin, USA, 2007.
(b) Bruker. SADABS; Bruker AXS Inc.: Madison, Wisconsin, USA. 2001. (c) Sheldrick, G.
M. Acta Cryst. 2008, A64, 112-122.

You might also like